ABSTRACT
OmpA, the most abundant porin in Stenotrophomonas maltophilia KJ, exists as a two-domain structure with an N-terminal domain of β-barrel structure embedded in the outer membrane and a C-terminal domain collocated in the periplasm. KJΔOmpA299-356, an ompA mutant of S. maltophilia KJ with a truncated OmpA devoid of 299 to 356 amino acids (aa), was able to stably embed in the outer membrane. KJΔOmpA299-356 was more susceptible to β-lactams than wild-type KJ. We aimed to elucidate the mechanism underlying the ΔompA299-356-mediated increase in β-lactam susceptibility (abbreviated as “ΔOmpA299-356 phenotype”). KJΔOmpA299-356 displayed a lower ceftazidime (CAZ)-induced β-lactamase activity than KJ. Furthermore, KJ2, a L1/L2 β-lactamases-null mutant, and KJ2ΔOmpA299-356, a KJ2 mutant with truncated OmpA devoid of299 to 356 aa, had comparable β-lactam susceptibility. Both lines of evidence indicate that decreased β-lactamase activity contributes to the ΔOmpA299-356 phenotype. We analyzed the transcriptome results of KJ and KJΔOmpA299-356, focusing on PG homeostasis-associated genes. Among the 36 genes analyzed, the nagA gene was upregulated 4.65-fold in KJΔOmpA299-356. Deletion of the nagA gene from the chromosome of KJΔOmpA299-356 restored β-lactam susceptibility and CAZ-induced β-lactamase activity to wild-type levels, verifying that nagA-upregulation in KJΔOmpA299-356 contributes to the ΔOmpA299-356 phenotype. Furthermore, transcriptome analysis revealed that rpoE (Smlt3555) and rpoP (Smlt3514) were significantly upregulated in KJΔOmpA299-356. The deletion mutant construction, β-lactam susceptibility, and β-lactamase activity analysis demonstrated that σP, but not σE, was involved in the ΔOmpA299-356 phenotype. A real-time quantitative (qRT-PCR) assay confirmed that nagA is a member of the σP regulon. The involvement of the σP-NagA-L1/L2 regulatory circuit in the ΔOmpA299-356 phenotype was manifested.
IMPORTANCE Porins of Gram-negative bacteria generally act as channels that allow the entry or extrusion of molecules. Moreover, the structural role of porins in stabilizing the outer membrane by interacting with peptidoglycan (PG) and the outer membrane has been proposed. The linkage between porin deficiency and antibiotic resistance increase has been reported widely, with a rationale for blocking antibiotic influx. In this study, a link between porin defects and β-lactam susceptibility increase was demonstrated. The underlying mechanism revealed that a novel σP-NagA-L1/L2 regulatory circuit is triggered due to the loss of the OmpA-PG interaction. This study extends the understanding on the porin defect and antibiotic susceptibility. Porin defects may cause opposite impacts on antibiotic susceptibility, which is dependent on the involvement of the defect. Blocking the porin channel role can increase antibiotic resistance; in contrast, the loss of porin structure role may increase antibiotic susceptibility.
KEYWORDS: OmpA, sigma factor, beta-lactam resistance, peptiodglycan stress
INTRODUCTION
The outer membrane, a unique organelle of Gram-negative bacteria, protects them against harsh environments (1). Porins are the most abundant proteins in the outer membrane. They are designed for exchanging molecules across the outer membrane (2) and act as crucial factors in cell-to-cell signaling and environmental sensing. Porins are classified into the following two types according to their physiological roles: classical and slow porins (3). Classical porins serve as the entry point for molecules in a nonselective fashion (such as OmpF and OmpC of Escherichia coli and OmpK36 of Klebsiella pneumonia) (3, 4) or in a substrate-specific manner (such as LamB of E. coli and ScrY of Salmonella enterica subsp. Typhimurium) (5, 6). Relative to classical porins, molecular transportation is not the major function of slow porins because of their very low permeability. A classic example of a slow porin is the OmpA. OmpA is an abundant β-barrel porin highly conserved among bacterial species and is characterized as a two-domain structure with an N-terminal domain of β-barrel structure embedded in the outer membrane and a C-terminal OmpA-like globular domain located in the periplasmic space (7–10). Critical biological functions of the periplasmic globular domain of OmpA have been revealed, including noncovalent association with the peptidoglycan (PG) layer (11–13), thereby maintaining outer membrane integrity and signal transduction. OmpA is a multifaceted outer membrane protein (OMP) that is involved in a number of functions, such as adhesion, invasion, swimming, serum resistance, biofilm formation, and antibiotic resistance (14, 15), and serves as a receptor for pilus, bacteriophages, and bacteriocins (16). Furthermore, OmpA is an immune target that induces a host immune response; this feature makes OmpA the most popular vaccine candidate against Gram-negative pathogens (17).
Bacterial cell walls are essential for bacterial viability because they provide structural strength and counteract osmotic pressure in the cytoplasm. PG, which comprises sugars and amino acids, is a critical component of the cell wall (18). Given its uniqueness to bacteria, PG is a promising target for antibiotics, such as β-lactams. The target of β-lactam action is penicillin-binding proteins (PBPs), which participate in PG synthesis (19).
The known mechanisms responsible for β-lactam resistance include decreased outer membrane permeability, overexpression of efflux pumps, mutation of β-lactam targets, and overexpression of β-lactamase (20). β-Lactamase overexpression remains the primary mechanism used by Gram-negative bacteria to withstand β-lactam action. In some Gram-negative bacteria (such as Enterobacter cloacae, Citrobacter freundii, Pseudomonas aeruginosa, and Stenotrophomonas maltophilia), the inducible expression of chromosomally encoded β-lactamase is tightly linked to the cell wall recycling process (21). The cleaved PG sacculus in the periplasm is transported into the cytosol and further processed into different derivatives. Of these derivatives, 1,6-anhydro-MurNAc-peptides and UPD-MurNAc-pentapeptides function as activator and repressor ligands, respectively, which bind with AmpR and induce and repress the expression of β-lactamase genes, respectively (22–24).
Bacterial gene expression is often regulated at the transcriptional level. RNA polymerase (RNAP), an enzyme complex responsible for transcription, is essential to life. Bacterial RNAP consists of six subunits, as follows: αI, αII, β, β’, ω, and σ (25). The transcription process begins with the assembly of αI, αII, β, β’, and ω subunits into a core RNAP complex, and then a σ factor is recruited to form the holoenzyme. Bacteria generally harbor several different σ factors that specifically switch gene expression (26). The sigma factor recognizes cognate promoter sequences upstream of the genes that comprise the regulon of the σ factor (27). Based on their sequence, domain architecture, and function, σ factors fall into two distinct families, as follows: the σ54 factor (RpoN) family and σ70 family (RpoD) (27). The σ70 family factors have been classified into the following four groups: I, II, III, and IV. Group IV σ factors (or extracytoplasmic function [ECF] σ factors) are capable of sensing and responding to signals generated outside the cell or in the cell envelope (28, 29). ECF factors participate in several biological functions, such as envelope stress response, cell wall stress response, oxidative stress response, and iron transport (30).
Stenotrophomonas maltophilia, an opportunistic pathogen, is increasingly being recognized as an important cause of nosocomial infections. S. maltophilia is intrinsically resistant to several antibiotics because it possesses a number of antibiotic resistance determinants, such as β-lactamases, efflux pumps, and aminoglycoside-modifying enzymes (31). Thus, the challenge in the treatment of S. maltophilia infection is increasing.
S. maltophilia is intrinsically resistant to most β-lactams because of chromosomally encoded L1 and L2 β-lactamases. Of the β-lactams, ceftazidime (CAZ) and ticarcillin-clavulanic acid (TIM) are the choices used for treating S. maltophilia infection. L1 and L2 β-lactamase-inducible expression of S. maltophilia is linked to the disturbance of PG homeostasis (32), such as AmpC expression in E. cloacae, C. freundii, and P. aeruginosa (22). In addition to β-lactamase, non-β-lactamase-mediated β-lactam resistance in S. maltophilia is also reported. For example, the loss of function of phoPQ results in the alteration of outer membrane permeability, which is involved in the compromise of β-lactam resistance (32).
OmpA (Smlt0955) of S. maltophilia is known to be the highly expressed gene in logarithmically grown S. maltophilia (15, 33). The relationship between ompA deletion, swimming compromise, and conjugation failure has been reported in our recent study (15). In this study, we aimed to elucidate the relationship between OmpA defects and antibiotic susceptibility of S. maltophilia. Our findings revealed the involvement of N-acetylglucosamine-6-phosphate deacetylase (NagA) and a novel ECF σ factor (σP) in the β-lactam susceptibility of S. maltophilia.
RESULTS
The truncated OmpA protein expressed in KJΔOmpA299-356 is embedded in the outer membrane.
KJΔOmpA, an in-frame ompA deletion mutant, was constructed in our previous study (15) in which the C-terminal OmpA c-like domain was partially deleted (299 to 356 amino acids) (Fig. 1A). For a more precise description of its characteristics, we renamed KJΔOmpA (15) as KJΔOmpA299-356 here. As the N-terminal β-barrel domain of OmpA remained intact in KJΔOmpA299-356, we wondered whether the truncated OmpA protein could be embedded in the outer membrane. The OMPs of KJ and KJΔOmpA299-356 were purified and subjected to sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) analysis. A comparison between the OMP profiles of both strains revealed that KJΔOmpA299-356 was short of a protein band (band A in Fig. 1B) and had two additional protein bands compared to wild-type KJ (bands B and C in Fig. 1B). The three protein bands were excised from the gel and characterized using liquid chromatography-tandem mass spectrometry (LC-MS/MS). The masses and fragmentation patterns of band A correlated with OmpA, proving the correctness of KJΔOmpA299-356 (see Table S1 in the supplemental material). As determined using LC-MS/MS, band B was also identified as OmpA. It is worth mentioning that the fragmentation patterns of band B covered the N-terminal β-barrel domain of OmpA but not the predicted signal peptide (1- to 22-amino acid residues) or the deleted C-terminal region (see Table S2 in the supplemental material). The expected truncated OmpA protein of KJΔOmpA299-356 should have a molecular weight of 31.1 kDa, which matches the location of band B in the gel (Fig. 1Β). The LC-MS/MS results for band C revealed Smlt4119 as a candidate (see Table S3 in the supplemental material). Smlt4119 is predicted to be a 272-aa OMP with a 25-aa signal peptide. The predicted molecular weight of mature Smlt4119 (without the signal peptide) was 27.9 kDa, consistent with its position in the SDS-PAGE gel (Fig. 1B). Collectively, our results supported that the truncated OmpA protein of KJΔOmpA299-356 was able to stably embed in the outer membrane (Fig. 1C).
FIG 1.
Construction strategy and outer membrane protein profiling of KJΔOmpA299-356, an ompA in-frame deletion mutant. (A) Diagram of conserved protein domains of OmpA and the deleted region of the ompA gene in KJΔOmpA299-356. The OMP β-barrel and OmpA c-like domains of the OmpA protein are marked in blue and red, respectively. The deleted region of the ompA gene is indicated as a white rectangle. (B) Outer membrane protein profiling of KJ and KJΔOmpA299-356. Outer membrane proteins were prepared from logarithmic-phase KJ and KJΔOmpA299-356 as described in the Materials and Methods and then separated by SDS-PAGE with a 5% stacking gel and a 15% separating gel. The gel was stained with 0.1% Coomassie brilliant blue and destained with 40% methanol-10% glacial acetic acid. Lane M, molecular weight standards; 1, KJ; 2, KJΔOmpA299-356. The black arrows indicate the protein excised for LC-MS/MS analysis. (C) An artistic impression of OmpA proteins in wild-type KJ and KJΔOmpA299-356 mutant.
KJΔOmpA299-356 is more susceptible to β-lactam than KJ.
Antibiotic susceptibilities of wild-type KJ and KJΔOmpA299-356 were examined. The antibiotics tested included β-lactam (ceftazidime [CAZ] and ticarcillin-clavulanic acid [TIM]), fluoroquinolone (ciprofloxacin, levofloxacin, and moxifloxacin), and trimethoprim-sulfamethoxazole, which are common choices for the treatment of S. maltophilia infection. Among the antibiotics tested, alteration in β-lactam susceptibility was the most apparent. Compared with wild-type KJ, KJΔOmpA299-356 reduced the MICs of TIM and CAZ (Table 1); however, the impact of truncated OmpA on susceptibility to quinolone and trimethoprim-sulfamethoxazole was mild and not significant (see Table S4 in the supplemental material).
TABLE 1.
Antibiotic susceptibilities of S. maltophilia KJ and its derived mutants
Strain | MIC (μg/mL) of:a |
|
---|---|---|
CAZ | TIM | |
KJ | 256 | 96 |
KJΔOmpA299-356 | 16 | 32 |
KJL2-OmpAΔOmpA299-356 | >256 | 192 |
KJ2 | 0.094 | 0.094 |
KJ2ΔOmpA299-356 | 0.094 | 0.125 |
KJΔNagAΔOmpA299-356 | >256 | 96 |
KJΔRpoEΔOmpA299-356 | 12 | 32 |
KJΔRpoPΔOmpA299-356 | 192 | 96 |
KJL2-RpoP | 48 | 48 |
KJL2-RpoNΔOmpA299-35 | 24 | 32 |
CAZ, ceftazidime; TIM, ticarcillin-clavulanate.
In our previous study, we demonstrated that KJΔOmpA299-356 is unable to obtain an exogenous plasmid by conjugation (15); thus, the plasmid-mediated complementation assay was inaccessible for KJΔOmpA299-356. Furthermore, from the transcriptome data (15), we noticed that the L2 transcript level in KJΔOmpA299-356 had a 3.82-fold increase compared with that in the wild-type KJ. Therefore, we constructed an alternative ompA complementary strain KJL2-OmpAΔOmpA299-356, in which the complemented ompA gene was inserted downstream of the L2 gene without disrupting the L2 gene (see Fig. S1 in the supplemental material). The ompA gene insertion had no impact on L2 expression, which was verified by qRT-PCR (data not shown). The susceptibilities of KJL2-OmpAΔOmpA299-356 to TIM and CAZ were examined, and reverted MIC values were observed (Table 1).
The known mechanisms responsible for the β-lactam resistance of S. maltophilia can be classified into β-lactamase-mediated resistance and non-β-lactamase-mediated resistance (33, 34). The β-lactamase activities of KJ, KJΔOmpA299-356, and KJL2-OmpAΔOmpA299-356 were determined to assess the involvement of β-lactamases in the ΔompA299-356-mediated increase in β-lactam susceptibility. Compared with wild-type KJ, KJΔOmpA299-356 had lower CAZ-induced β-lactamase activity and KJL2-OmpAΔOmpA299-356 exhibited comparable activities (Fig. 2). Next, non-β-lactamase-mediated resistance was further assessed in the L1 and L2 double deletion mutant, KJ2 (35). The ΔompA299-356 allele was introduced into the chromosome of KJ2 to yield KJ2ΔOmpA299-356. KJ2ΔOmpA299-356 and the parental strain KJ2 displayed comparable susceptibility to TIM and CAZ (Table 1). Collectively, decreased β-lactamase activity was the dominant factor responsible for the ΔompA299-356-mediated increase in β-lactam susceptibility.
FIG 2.
CAZ-induced β-lactamase activities of S. maltophilia KJ and its derived in-frame deletion constructs. Overnight-cultured bacterial cells were inoculated into fresh LB broth at an initial OD450 of 0.15 and subcultured for 3 h. Induction was carried out by adding CAZ of 1/4 MIC for 1 h, and the induced β-lactamase activities were determined. Bars represent the average values from three independent experiments. Error bars represent the SEM. *, P < 0.001; significance was calculated by Student’s t test. CAZ, ceftazidime.
NagA (Smlt4020) upregulation in KJΔOmpA299-356 is involved in the ΔompA299-356-mediated increase in β-lactam susceptibility.
The RNA sequencing (RNA-Seq) transcriptome assays of KJ and KJΔOmpA299-356 were conducted and validated in our previous study (15) (see Table S5 in the supplemental material). To elucidate the mechanism underlying the ΔompA299-356-mediated increase in β-lactam susceptibility, we reanalyzed the transcriptome data (15) (Table S5). It is commonly recognized that PG homeostasis is linked to the expression of chromosomally encoded β-lactamase genes in S. maltophilia (32); thus, genes associated with PG homeostasis were surveyed. The PG homeostasis model for S. maltophilia has been proposed in our previous study, and there are at least 39 genes involved (32). Based on this model, we re-examined the transcript levels of the 39 genes from previous transcriptome data (15) (Table S5). We defined significance as the absolute fold change in TPM equal to or greater than 3. Among the 36 genes examined, nagA (Smlt4020) was significantly upregulated (Fig. 3; see Table S6 in the supplemental material). The expression of nagA in KJ and KJΔOmpA299-356 was validated using qRT-PCR (Fig. 4). The protein encoded by Smlt4020 is annotated as N-acetylglucosamine-6-phosphate deacetylase (NagA), which catalyzes the deacetylation of N-acetylglucosamine-6-phosphate (GlcNAc-6P) to glucosamine-6-phosphate (GlcN-6P). GlcN-6P is then used in two main pathways in bacteria, as follows: the PG recycling pathway and glycolysis pathway (36).
FIG 3.
Schematic representation of PG biosynthesis, catabolism, and recycling, and the comparative transcriptome analysis in S. maltophilia KJ and its derived ompA mutant, KJΔOmpA299-356. The PG homeostasis model of S. maltophilia is proposed based on the known model from P. aeruginosa. The PG biosynthesis (labeled in purple) and PG catabolism (labeled in blue) majorly occur in the periplasm. AmpNG permease system transports the PG degradation fragments from periplasm into cytosol for further PG recycling (labeled in green). The dashed lines indicate the homologs of P. aeruginosa are not identified from S. maltophilia K279a genome via BLAST analysis. The RNA-Seq transcriptome analysis of KJ and KJΔOmpA299-356 was performed. The number in each bracket indicates the fold of gene expression change in wild-type KJ and KJΔOmpA299-356. Positive values indicate the gene expression in KJΔOmpA299-356 is higher than that in wild-type KJ, and negative values represent the gene expression in KJ is higher than that in KJΔOmpA299-356.
FIG 4.
The nagA transcript level of wild-type KJ and its derived mutants. Overnight culture of strains tested were inoculated into fresh LB at an initial OD450 of 0.15 and aerobically grown for 5 h. The nagA transcript level was determined via qRT-PCR using NagAQ103-F/R primers. Relative transcript level was normalized to the transcript level of KJ cells. Bars represent the average values from three independent experiments. Error bars represent the standard error of the mean. *, P < 0.05; significance was calculated by Student’s t test.
To clarify whether nagA upregulation in KJΔOmpA299-356 was responsible for the ΔompA299-356-mediated increase in β-lactam susceptibility, a nagA deletion mutation was introduced into KJΔOmpA299-356. Introduction of the ΔnagA allele into KJΔOmpA299-356 restored β-lactam resistance (Table 1) and β-lactamase activity (Fig. 2), supporting that nagA overexpression in KJΔOmpA299-356 contributes to ΔompA299-356-mediated increase in β-lactam susceptibility.
RpoP (Smlt3514) upregulation in KJΔOmpA299-356 is involved in the ΔompA299-356-mediated increase in β-lactam susceptibility.
Given that OmpA is the most abundant OMP in wild-type KJ, we speculated that ompA deletion might affect outer membrane integrity. Thus, the outer membrane destabilization of KJΔOmpA299-356 was investigated using the 1-N-phenylnaphtylamine (NPN) uptake assay. NPN is an uncharged lipophilic dye with weak fluorescence in aqueous environments and great fluorescence in hydrophobic environments, such as the cell membrane. If the outer membrane integrity of KJΔOmpA299-356 is compromised, NPN dye is integrated into the inner membrane. The NPN assay demonstrated that the level of fluorescence intensity of KJΔOmpA299-356 was 1.95-fold higher than that of wild-type KJ and nearly reverted to the wild-type level in the complementary strain KJL2-OmpAΔOmpA299-356 (Fig. 5). It was worth mentioning that the fluorescence detected from KJΔNagAΔOmpA299-356 was comparable to that from KJΔOmpA299-356 (Fig. 5), indicating that nagA deletion from KJΔOmpA299-356 hardly reverted the outer membrane defect.
FIG 5.
1-N-Phenylnaphthylamine (NPN) uptake assay. The 100-μL bacterial cells tested were pipetted into 96-well microtiter plates, and NPN was added to a final concentration of 15 μM. Fluorescence was monitored by fluorescence spectrophotometer at excitation and emission wavelengths of 355 nm and 402 nm, respectively. Relative fluorescence level was normalized to KJ cells. Bars represent the average values from three independent experiments. Error bars represent the standard error of the mean. *, P < 0.05; significance was calculated by Student’s t test.
Extracytoplasmic function (ECF) σ factors provide a means of regulating gene expression in response to extracytoplasmic stress, such as imbalance of the outer membrane, peptidoglycan, and inner membrane (37). Given that the membrane integrity was compromised in KJΔOmpA299-356, the involvement of the ECF σ factor was considered. There were 16 annotated ECF σ factors in the S. maltophilia K279a genome (38). Transcriptome profiles of the 16 ECF σ factors in wild-type KJ and KJΔOmpA299-356 (15) (Table S5) were investigated. Among the 16 ECF σ factors assayed, two sigma factors, namely, Smlt3514 and rpoE, displayed significant expression differences with 3.15- and 5.58-fold upregulation in KJΔOmpA299-356, respectively (Table 2). Based on the following results, we designated Smlt3514 as rpoP (P meaning PG) here.
TABLE 2.
Transcriptome analysis of ECF σ factor genes in wild-type KJ and ompA mutant, KJΔOmpA299-356
Locus | Protein | TPMa of |
Fold changeb | |
---|---|---|---|---|
KJ | KJΔOmpA299-356 | |||
Smlt0855 | ECF σ factor | 14.17 | 19.88 | +1.40 |
Smlt1269 | ECF σ factor | 9.79 | 12.56 | +1.28 |
Smlt1349 | FecI-like σ factor | 85.82 | 124.59 | +1.45 |
Smlt1750 | FecI-like σ factor | 4.81 | 8.64 | +1.80 |
Smlt2004 | ECF σ factor | 79.35 | 57.35 | −1.38 |
Smlt2377 | ECF σ factor | 63.83 | 23.17 | −2.75 |
Smlt2513 | ECF σ factor | 1.44 | 1.08 | −1.34 |
Smlt2664 | FecI-like σ factor | 8.25 | 3.21 | −2.57 |
Smlt2716 | FecI-like σ factor | 17.87 | 14.17 | −1.26 |
Smlt2848 | FecI-like σ factor | 0 | 0 | ND |
Smlt2935 | ECF σ factor | 0.99 | 0 | ND |
Smlt3223 | ECF σ factor | 2.47 | 2.23 | −1.11 |
Smlt3514 | σP | 49.65 | 156.39 | +3.15 |
Smlt3555 | σE | 434.09 | 2,424.29 | +5.58 |
Smlt3900 | ECF σ factor | 11.04 | 11.05 | 1.00 |
Smlt4579 | ECF σ factor | 0 | 0 | ND |
TPM, transcripts per kilobase million.
Negative fold changes represent genes that were significantly downregulated in KJΔOmpA299-356, whereas positive fold changes represent upregulation in KJΔOmpA299-356. ND, not determined.
The Smlt3514 gene encodes a 192-aa sigma-70 family, ECF subfamily RNA polymerase sigma factor. A survey on the 14 sequenced S. maltophilia strains revealed that this gene is completely conserved in this species, with an intraspecies protein identity of 93 to 100%. To clarify the relationship between σP, σE, and ΔompA299-356-mediated increase in β-lactam susceptibility, we constructed rpoE and rpoP in-frame deletion mutants of KJΔOmpA299-356 (KJΔRpoEΔOmpA299-356 and KJΔRpoPΔOmpA299-356) and examined the β-lactam susceptibility and β-lactamase activity of these mutants. Compared with KJΔOmpA299-356, KJΔRpoPΔOmpA299-356, but not KJΔRpoEΔOmpA299-356, almost restored β-lactam resistance and β-lactamase activity to wild-type levels (Table 1, Fig. 2).
NagA upregulation in KJΔOmpA299-356 is σP dependent.
The next question we wondered was whether nagA is a member of the σP regulon. Thus, the nagA transcript level in KJΔRpoPΔOmpA299-356 was determined by qRT-PCR. Compared with that in wild-type KJ, the nagA transcript level of KJΔOmpA299-356 showed a 3.52- ± 0.49-fold increase and reverted to the wild-type level when ΔrpoP allele was introduced into the chromosome of KJΔOmpA299-356 (Fig. 4).
Based on the above results, we conclude that the ΔompA299-356-mediated increase in β-lactam susceptibility seems to be attributed to the σP-NagA-L1/L2 regulatory circuit. To further confirm this possibility, we constructed an rpoP-overexpression strain, KJL2-RpoP, in which the rpoP gene was inserted downstream of the L2 gene (Fig. S1). Compared with wild-type KJ, KJL2-RpoP was more susceptible to β-lactams (Table 1) and had a decreased CAZ-induced β-lactamase activity (Fig. 2).
DISCUSSION
Outer membrane porins are important channels for the influx of nutrients, hydrophilic molecules, and some antibiotics. The linkage between porin deficiency and β-lactam resistance has been reported widely; some examples are OmpK35 and OmpK36 of Klebsiella pneumoniae (39), OmpF of Serratia marcescens (40), and OprD and OprF of P. aeruginosa (41, 42). These porins (OmpK, OmpF, OprD, or OprF) are classical ones that generally function as outlets for molecule entrance or extrusion (43), and inactivation of these porins generally results in an increase in antibiotic resistance. Nevertheless, the OmpA is a slow porin. Smani et al. (44) found that ompA inactivation of A. baumannii decreased the β-lactam MICs and proposed that OmpA participates in antimicrobial extrusion, but they did not verify this assumption. The rationale proposed by Smani et al. (44) was based on the concept of OmpA as a transport channel for antibiotics. In this study, we revealed that the loss of function of OmpA of S. maltophilia compromised β-lactam resistance and disclosed the involvement of the σp-NagA-L1/L2 regulatory circuit in this phenotype. Our findings emphasize the role of OmpA in stabilizing the PG layer, which is not limited to molecule transport. Disruption of PG homeostasis causes extracytoplasmic stress, which leads to ECF sigma factor activation and increases β-lactam susceptibility.
OmpA adopts a two-domain structure, with an N-terminal β-barrel embedded in the outer membrane (8) and a C-terminal globular domain in the periplasm (9, 10). Critical biological functions of the periplasmic C-terminal domain of OmpA have been revealed, including noncovalent association with the PG scaffold (12) and in complex with the RcsF lipoprotein (13). The stable binding of PG to the C-terminal domain of OmpA is a recognition mechanism for Gram-negative bacteria to maintain cell wall integrity (12). Structural analysis of Acinetobacter baumannii OmpA clearly indicated that Asp271 and Arg286 are key residues for OmpA and PG interactions (10). By protein alignment between OmpAs of A. baumannii and S. maltophilia, we found that the two residues are well conserved in S. maltophilia OmpA (Asp294 and Arg309) (see Fig. S2 in the supplemental material) and that Arg309 was deleted in KJΔOmpA299-356. Thus, it is not surprising that a loss of the C-terminal domain of OmpA may cause cell wall and/or cell envelope stress, which then triggers the activation of an ECF σ factor. As revealed in this study, KJΔOmpA299-356 exerted an effect on the upregulation of rpoP. The stimulus that triggers the upregulation of rpoP can be a compromise of PG stability caused by the loss of interaction between PG and the OmpA C-terminal domain, which can be regarded as a type of PG stress. This is the reason why we designate Smlt3514 as rpoP. Since the ompA mutant constructed in this study is a partial OmpA deletion mutant with an intact β-barrel component, we cannot immediately rule out the possibility that the β-barrel domain of OmpA is also required for the activation of σP, in addition to the loss of the C-terminal domain.
The SDS-PAGE OMP profiling revealed that an OMP encoded by Smlt4119 was invisible in wild-type KJ but obviously increased in KJΔOmpA299-356 (Fig. 1B). The Smlt4119 protein is annotated as a hypothetical protein in the sequenced S. maltophilia genome and is not homologous to other known OMPs. The question of whether the increased Smlt4119 protein level has an impact on the ΔompA299-356-mediated increase in β-lactam susceptibility is not immediately clear right now. However, KJΔNagAΔOmpA299-356 and KJΔRpoPΔOmpA299-356 displayed comparable β-lactam susceptibility to wild-type KJ (Table 1), highly supporting that σP-NagA-L1/L2 is the major regulatory circuit involved in the ΔompA299-356-mediated increase in β-lactam susceptibility.
Recently, we have demonstrated that rpoN was downregulated in KJΔOmpA299-356, which results in the swimming compromise of KJΔOmpA299-356 (15). In this article, we further elucidated that rpoP upregulation in KJΔOmpA299-356 contributes to a Δomp299-356-mediated increase of β-lactam susceptibility. We were curious whether an interconnection between σN- and σP-mediated regulations happens. First, the β-lactam susceptibility of KJL2-RpoNΔOmpA299-356, a rpoN complementation strain of ΔOmpA299-356, was investigated. KJL2-RpoNΔOmpA299-356 exhibited comparable β-lactam susceptibility to KJΔOmpA299-356 (Table 1), indicating that rpoN downregulation in KJΔOmpA299-356 is less related to the ΔompA299-356-mediated increase of β-lactam susceptibility. In addition, KJΔRpoPΔOmpA299-356 and KJΔOmpA299-356 displayed comparable swimming motility (see Fig. S3 in the supplemental material), which is less support for the involvement of σP in ΔompA299-356-mediated swimming compromise. Collectively, σN- and σP-mediated regulations in KJΔOmpA299-356 appeared to independently link to the phenotypes of swimming compromise and β-lactam susceptibility increase.
PG recycling is a process whereby degraded PGs are recovered and made available for bacterial cells to synthesize more PG or to use it as an energy source. NagA is an enzyme involved in PG recycling (Fig. 3). NagA catalyzes the conversion of GlcNAc-6P to GlcN-6P, and GlcN-6P is used in PG recycling or glycolysis pathways. In the PG recycling pathway, GlcN-6P is subsequently processed into UPD-MurNAc-pentapeptide, which acts as a repressor ligand to bind with AmpR. UPD-MurNAc-pentapeptide-bound AmpR functions as a repressor of chromosomal β-lactamase genes expression (22). Based on this rationale, we propose an explanation for the role of nagA in the ΔompA-mediated increase in β-lactam susceptibility. The nagA upregulation in KJΔOmpA299-356 may increase the intracellular levels of UPD-MurNAc-pentapeptides, which attenuates the association between 1,6-anhydro-MurNAc-peptides and AmpR. This finding may explain why the β-lactamase activity of KJΔOmpA299-356 decreased.
The linkage between σ factor and β-lactam resistance has been reported in several bacteria, for example σP in Bacillus anthracis, Bacillus cereus, and Bacillus thuringiensis (45, 46); σM, σX, and σA in Bacillus subtilis (47, 48); σB in Staphylococcus aureus (49); algT/U in Pseudomonas aeruginosa (50); ECF-10 in Pseudomonas putida (51); and σE7 and σH3 in Azospirillum baldaniorum (52). The σP of S. maltophilia displayed 25 to 27% and 46 to 54% protein identity and similarity, respectively, to the σP of Bacillus spp. Some sigma factors, such as σP and σM, are modulated by anti-σP via protein-protein interactions (45, 46). These σ factors are generally held inactive by anti-σ factors, and the genes encoding σ factor and anti-σ factor are organized into an operon (53, 54). We surveyed the genomic organization surrounding rpoP, but none of the open reading frames (ORFs) exhibited any homology to the anti-σ factor. However, interestingly, we noticed that the genes (Smlt3513 and Smlt3512) downstream of rpoP encode a 177-aa and a 293-aa hypothetical protein, respectively. The rpoP, Smlt3513, and Smlt3512 genes have the same orientation. RpoP and Smlt3513 genes have a 4-bp overlapping, and Smlt3513 and Smlt3512 genes are separated by a 10-bp intergenic region. Based on this observation, we speculate that rpoP-Smlt3513-Smlt3512 may form an operon and Smlt3513 and Smlt3512 may participate in the PG stress signal transduction from the periplasm into cytoplasm and then activate the cytoplasmic σP.
The genetic indications presented here lead to a model in which a σP-NagA-L1/L2 regulatory circuit is responsible for the increase in β-lactam susceptibility in KJΔOmpA299-356. OmpA has a two-domain structure consisting of a β-stranded N-terminal domain and a globular C-terminal domain (16). The stable interaction between the C-terminal domain of OmpA and PG is critical for the maintenance of cell wall integrity (12) (Fig. 6A). When S. maltophilia KJ is challenged with CAZ, PG homeostasis is disturbed and abundant 1,6-anhydro-MurNAc-pentapeptides (activator ligands) dominate the binding with AmpR, which leads to L1/L2 β-lactamases upregulation and enhances β-lactam resistance (Fig. 6A). In the KJΔOmpA229-356 mutant, the OmpA protein lacks 229- to 356-amino acid residues but retains an intact N-terminal β-barrel domain, which allows it to assemble into the outer membrane. Due to the deletion of the C-terminal domain, OmpA loses its interaction with PG layers, which destroys PG stability and generates PG stress. In response to PG stress, ECF sigma factor σP is upregulated, which increases the expression of NagA. High expression of NagA may favor the formation of UPD-MurNAc-pentapeptides (repressor ligands). In such an instance, increased repressor ligands would alter the repressor ligand/activator ligand ratio, decrease the association of activator ligands and AmpR, and result in the reduction of β-lactam-induced β-lactamase activity (Fig. 6B).
FIG 6.
Model of the rpoP-nagA-L1/L2 regulatory circuit in the ΔompA299-356-mediated increment of β-lactam susceptibility of Stenotrophomonas maltophilia. (A) The PG scaffold of Gram-negative bacteria is anchored noncovalently to the outer membrane via the C-terminal domain of OmpA proteins. The OmpA-PG interaction controls the stability of the cell wall. Ceftazidime (CAZ), a β-lactam antibiotic, targets on the penicillin binding protein (PBP) and blocks PG cross-linking. Accumulated murein sacculus is transported into the cytosol via the AmpNG permease system and further processed into 1,6-anhydro-MurNAc-pentapeptides (activator ligands). The activator ligands compete with UDP-MurNAc-pentapeptides (repressor ligands) for a binding site on AmpR. As AmpR is bound with activator ligands, it functions as a transcriptional activator, inducing the expression of L1 and L2 β-lactamases (L2 as a representative). (B) In the KJΔOmpA229-356 mutant, the C terminus-deleted OmpA loses the interaction with PG layers, generating a PG stress. PG stress triggers the upregulation of σP, which then increases the expression of nagA. The upregulated NagA activity may favor the formation of repressor ligands. The increased repressor ligands are though to partially displace activator ligands from AmpR and attenuate CAZ-induced β-lactamases expression.
MATERIALS AND METHODS
Bacterial strains, plasmids, and primers.
The bacterial strains, plasmids, and primers used in this study were summarized in Table S7 in the supplemental material.
OMP preparation and SDS-PAGE.
The outer membrane proteins of the mid-log-phase bacterial cells were prepared as described previously (55). The OMPs were separated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) in a 15% polyacrylamide gel. The gel was stained using 0.1% Coomassie brilliant blue R250 (Bio-Rad) and destained with 40% methanol-10% glacial (acetic acid) for the visualization of proteins.
Construction of deletion mutants.
The deletion mutants were constructed using double crossover homologous recombination as described previously (56). Two DNA fragments flanking the rpoE were PCR amplified using primer sets RpoEN-F/RpoEN-R and RpoEC-F/RpoEC-R (Table S7) and then subsequently cloned into pEX18Tc to generate pΔRpoE (Table S7). The intact nagA and rpoP genes were amplified by PCR using primer sets NagA-F/NagA-R and RpoP-F/RpoP-R (Table S7) and then cloned into pEX18Tc, yielding pEXNagA and pEXRpoP, respectively. The pΔNagA and pΔRpoP (Table S7) were generated by removing a 432-bp PstI-PstI and a 369-SalI-SlaI DNA fragments from plasmid pEXNagA and pEXRpoP, respectively. Plasmids pΔRpoE, pΔNagA, and pΔRpoP were transferred into S. maltophilia KJ by conjugation. The plasmid’s conjugation, transconjugant’s selection, and mutant’s confirmation were carried out as described previously (56).
Construction of KJL2-OmpAΔOmpA299-356 and KJL2-RpoP.
As the conjugation for the plasmid transportation is inapplicable in KJΔOmpA299-356 (15), an alternative strategy was designed for gene expression in KJΔOmpA299-356 by chromosomally inserting the ompA and rpoP genes, respectively, downstream L2 gene without disrupting any gene. The L2 and the inserted gene form an operon-like configuration, and the inserted gene is expressed inducibly upon β-lactam challenge (Fig. S1). The 503- and 547-bp DNA fragments containing the C terminus of the L2 gene and downstream of the L2 gene were obtained by PCR using the primer sets of HH1N-F/HH1N-R and HH1C-F/HH1C-R (Table S7), respectively, and then subsequently cloned into pEX18Tc, yielding plasmid pEXHH1 (Table S7). The multiple cloning sites (SphI/PstI/SalI/XbaI/BamHI/SmaI/KpnI/SacI) of pEX18Tc were conserved in plasmid pEXHH1 for cloning the exotic gene intended to be expressed. The intact ompA and rpoP genes were amplified by PCR using the primers sets HHOmpA-F/HHOmpA-R and HHRpoP-F/HHRpoP-R (Table S7) and then cloned into pEXHH to generate pEXHH1-OmpA and pEXHH1-RpoP (Table S7), respectively. The ompA and rpoP genes in pEXHH1-OmpA and pEXHH1-RpoP were inserted onto the intergenic region (IG) downstream the L2 gene of S. maltophilia KJ via double crossover homologous recombination as described previously (56) to yield KJL2-OmpA and KJL2-RpoP, respectively. The prototype chromosomal ompA gene was deleted from KJL2-OmpA by double crossover homologous recombination, and KJL2-OmpAΔOmpA299-356 was obtained.
Antibiotic susceptibility test.
The bacterial susceptibilities to antibiotics were determined by the Etest strips (bioMérieux, Marcy I’Etoile, France), according to the manufacturer’s instructions.
β-Lactamase activity determination.
The β-lactamase activity was determined using the chromogenic substrate nitrocefin (Δε = 20,500 M−1 · cm−1 at 486 nm) as the substrate as described previously (56). One unit of enzyme activity (U) was defined as the amount of enzyme that converts 1 nmol nitrocefin per minute. Specific activity (U/mg) was expressed as nanomoles of nitrocefin hydrolyzed per minute per milligram of protein.
Real-time quantitative PCR (qRT-PCR).
The preparation of DNA-free RNA and reverse transcription were carried out as described previously (35). qRT-PCRs were performed using the ABI StepOnePlus real-time PCR system. The primers used for qRT-PCR are listed in Table S7. The relative expression levels were determined by normalizing transcription to 16S rRNA and calculated using the threshold cycle (ΔΔCT) method (57). Each assay was performed at least three times by independent experiments.
N-phenylnaphtylamine (NPN) uptake assay.
The logarithmically grown cells were harvested, washed with 5 mM HEPES buffer (pH 7.2), and adjusted to an optical density at 450 nm (OD450) of 0.5 using the same buffer. The 96-well microtiter plate wells were supplemented with 100-μL bacterial suspensions and 15 μM NPN. After a 5-min incubation, the fluorescence was monitored using a fluorescence spectrophotometer (Tecan Infinite 200 PRO) at excitation and emission wavelengths of 355 nm and 402 nm, respectively. Fluorescence is emitted by NPN only after it partitions into the membrane; therefore, a greater emission of fluorescence represents greater outer membrane permeability to NPN.
Data availability.
The RNA-seq data have been deposited in GenBank BioSample accessions SAMN25290492 for S. maltophilia KJ and SAMN30672914 for KJΔOmpA299-356.
Footnotes
Supplemental material is available online only.
Contributor Information
Tsuey-Ching Yang, Email: tcyang@nycu.edu.tw.
Silvia T. Cardona, University of Manitoba
REFERENCES
- 1.Silhavy TJ, Kahne D, Walker S. 2010. The bacterial cell envelope. Cold Spring Harb Perspect Biol 2:a000414. doi: 10.1101/cshperspect.a000414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Koebnik R, Locher KP, Van Gelder P. 2000. Structure and function of bacterial outer membrane proteins: barrels in a nutshell. Mol Microbiol 37:239–253. doi: 10.1046/j.1365-2958.2000.01983.x. [DOI] [PubMed] [Google Scholar]
- 3.Nikaido H. 2003. Molecular basis of bacterial outer membrane permeability revisited. Microbiol Mol Biol Rev 67:593–656. doi: 10.1128/MMBR.67.4.593-656.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Dutzler R, Rummel G, Alberti S, Hernandez-Alles S, Phale P, Rosenbusch J, Benedi V, Schirmer T. 1999. Crystal structure and functional characterization of OmpK36, the osmotporin of Klebsiella pneumoniae. Structure 7:425–434. doi: 10.1016/s0969-2126(99)80055-0. [DOI] [PubMed] [Google Scholar]
- 5.Benz R, Schmid A, Nakae T, Vos-Scheperkeuter GH. 1986. Pore formation by LamB of Escherichia coli in lipid bilayer membranes. J Bacteriol 165:978–986. doi: 10.1128/jb.165.3.978-986.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Forst D, Welte W, Wacker T, Diederichs K. 1998. Structure of the sucrose-specific porin ScrY from Salmonella typhimurium and its complex with sucrose. Nat Struct Biol 5:37–46. doi: 10.1038/nsb0198-37. [DOI] [PubMed] [Google Scholar]
- 7.Vogel H, Jahnig F. 1986. Models for the structure of outer-membrane proteins of Escherichia coli derived from raman spectroscopy and prediction methods. J Mol Biol 190:191–199. doi: 10.1016/0022-2836(86)90292-5. [DOI] [PubMed] [Google Scholar]
- 8.Pautsch A, Schulz GE. 1998. Structure of the outer membrane protein A transmembrane domain. Nat Struct Biol 5:1013–1017. doi: 10.1038/2983. [DOI] [PubMed] [Google Scholar]
- 9.Mot R, Vanderleyden J. 1994. The C-terminal sequence conservation between OmpA-related outer membrane proteins and MotB suggests a common function in both Gram-positive and Gram-negative bacteria, possibly in the interaction of these domains with peptidoglycan. Mol Microbiol 12:333–334. doi: 10.1111/j.1365-2958.1994.tb01021.x. [DOI] [PubMed] [Google Scholar]
- 10.Park JS, Lee WC, Yeo KJ, Ryu KS, Kumarasiri M, Hesek D, Lee M, Mobashery S, Song JH, Kim SI, Lee JC, Cheong C, Jeon YH, Kim HY. 2012. Mechanism of anchoring of OmpA protein to the cell wall peptidoglycan of the gram-negative bacterial outer membrane. FASEB J 26:219–228. doi: 10.1096/fj.11-188425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Jahangiri A, Rasooli I, Owlia P, Fooladi AA, Salimian J. 2017. In silico design of an immunogen against Acinetobacter baumannii based on a novel model for native structure of outer membrane protein A. Microb Pathog 105:201–210. doi: 10.1016/j.micpath.2017.02.028. [DOI] [PubMed] [Google Scholar]
- 12.Samsudin F, Ortiz-Suarez ML, Piggot TJ, Bond PJ, Khalid S. 2016. OmpA: a flexible clamp for bacterial cell wall attachment. Structure 24:2227–2235. doi: 10.1016/j.str.2016.10.009. [DOI] [PubMed] [Google Scholar]
- 13.Dekoninck K, Létoquart J, Laguri C, Demange P, Bevernaegie R, Simorre JP, Dehu O, Iorga BI, Elias B, Cho SH, Collet JF. 2020. Defining the function of OmpA in the Rcs stress response. Elife 9:e60861. doi: 10.7554/eLife.60861. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Confer AW, Ayalew S. 2013. The OmpA family of proteins: roles in bacterial pathogenesis and immunity. Vet Microbiol 163:207–222. doi: 10.1016/j.vetmic.2012.08.019. [DOI] [PubMed] [Google Scholar]
- 15.Liao CH, Chang CL, Huang HH, Lin YT, Li LH, Yang TC. 2021. Interplay between OmpA and RpoN regulates flagellar synthesis in Stenotrophomonas maltophilia. Microorganisms 9:1216. doi: 10.3390/microorganisms9061216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Smith SG, Mahon V, Lambert MA, Fagan RP. 2007. A molecular Swiss army knife: OmpA structure, function and expression. FEMS Microbiol Lett 273:1–11. doi: 10.1111/j.1574-6968.2007.00778.x. [DOI] [PubMed] [Google Scholar]
- 17.Guan Q, Bhowmick B, Upadhyay A, Han Q. 2021. Structure and functions of bacterial outer membrane protein A, a potential therapeutic target for bacterial infection. Curr Top Med Chem 21:1129–1138. doi: 10.2174/1568026621666210705164319. [DOI] [PubMed] [Google Scholar]
- 18.Schleifer KH, Kandler O. 1972. Peptidoglycan types of bacterial cell walls and their taxonomic implications. Bacteriol Rev 36:407–477. doi: 10.1128/br.36.4.407-477.1972. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Lima LM, Silva BNMD, Barbosa G, Barreiro EJ. 2020. β-Lactam antibiotics: an overview from a medicinal chemistry perspective. Eur J Med Chem 208:112829. doi: 10.1016/j.ejmech.2020.112829. [DOI] [PubMed] [Google Scholar]
- 20.De Angelis G, Del Giacomo P, Posteraro B, Sanguinetti M, Tumbarello M. 2020. Molecular mechanisms, epidemiology, and clinical importance of β-lactam resistance in Enterobacteriaceae. Int J Mol Sci 21:5090. doi: 10.3390/ijms21145090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Dik DA, Fisher JF, Mobashery S. 2018. Cell-wall recycling of the Gram-negative bacteria and the nexus to antibiotic resistance. Chem Rev 118:5952–5984. doi: 10.1021/acs.chemrev.8b00277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Jacoby GA. 2009. AmpC beta-lactamases. Clin Microbiol Rev 22:161–182. doi: 10.1128/CMR.00036-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Johnson JW, Fisher JF, Mobashery S. 2013. Bacterial cell-wall recycling. Ann N Y Acad Sci 1277:54–75. doi: 10.1111/j.1749-6632.2012.06813.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Glen KA, Lamont IL. 2021. Beta-lactam resistance in Pseudomonas aeruginosa: current status, future prospects. Pathogens 10:1638. doi: 10.3390/pathogens10121638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Murakami KS. 2015. Structural biology of bacterial RNA polymerase. Biomolecules 5:848–864. doi: 10.3390/biom5020848. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Han L, Chen Q, Lin Q, Cheng J, Zhou L, Liu Z, Guo J, Zhang L, Cui W, Zhou Z. 2020. Realization of robust and precise regulation of gene expression by multiple sigma recognizable artificial promoters. Front Bioeng Biotechnol 8:92. doi: 10.3389/fbioe.2020.00092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Feklistov A. 2013. RNA polymerase: in search of promoters. Ann N Y Acad Sci 1293:25–32. doi: 10.1111/nyas.12197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Lonetto MA, Brown KL, Rudd KE, Buttner MJ. 1994. Analysis of the Streptomyces coelicolor sigE gene reveals the existence of a subfamily of eubacteria RNA polymerase sigma factors involved in the regulation of extracytoplasmic functions. Proc Natl Acad Sci USA 91:7573–7577. doi: 10.1073/pnas.91.16.7573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Helmann JD. 2019. Where to begin? Sigma factors and the selectivity of transcription initiation in bacteria. Mol Microbiol 112:335–347. doi: 10.1111/mmi.14309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Lonetto MA, Donohue TJ, Gross CA, Buttner MJ. 2019. Discovery of the extracytoplasmic function σ factors. Mol Microbiol 112:348–355. doi: 10.1111/mmi.14307. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Sánchez MB. 2015. Antibiotic resistance in the opportunistic pathogen Stenotrophomonas maltophilia. Front Microbiol 6:658. doi: 10.3389/fmicb.2015.00658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Huang HH, Wu BK, Li LH, Lin YT, Yang TC. 2021. Role of the PhoPQ two-component regulatory system in the β-lactam resistance of Stenotrophomonas maltophilia. J Antimicrob Chemother 76:1480–1486. doi: 10.1093/jac/dkab059. [DOI] [PubMed] [Google Scholar]
- 33.Chen YY, Wu HC, Lin JW, Weng SF. 2015. Functional properties of the major outer membrane protein in Stenotrophomonas maltophilia. J Microbiol 53:535–543. doi: 10.1007/s12275-015-5202-5. [DOI] [PubMed] [Google Scholar]
- 34.Wang Y, He T, Shen Z, Wu C. 2018. Antimicrobial resistance in Stenotrophomonas spp. Microbiol Spectr 6:6.1.04. doi: 10.1128/microbiolspec.ARBA-0005-2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Chen CH, Huang CC, Chung TC, Hu RM, Huang YW, Yang TC. 2011. Contribution of resistance nodulation-division efflux pump operon smeU1-V-W-U2-X to multidrug resistance of Stenotrophomonas maltophilia. Antimicrob Agents Chemother 55:5826–5833. doi: 10.1128/AAC.00317-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Park JT, Uehara T. 2008. How bacteria consume their own exoskeletons (turnover and recycling of cell wall peptidoglycan). Microbiol Mol Biol Rev 72:211–227. doi: 10.1128/MMBR.00027-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Raivio TL. 2005. Envelope stress response and Gram-negative bacterial pathogenesis. Mol Microbiol 56:1119–1128. doi: 10.1111/j.1365-2958.2005.04625.x. [DOI] [PubMed] [Google Scholar]
- 38.Crossman LC, Gould VC, Dow JM, Vernikos GS, Okazaki A, Sebaihia M, Saunders D, Arrowsmith C, Carver T, Peters N, Adlem E, Kerhornou A, Lord A, Murphy L, Seeger K, Squares R, Rutter S, Quail MA, Rajandream MA, Harris D, Churcher C, Bentley SD, Parkhill J, Thomson NR, Avison MB. 2008. The complete genome, comparative and functional analysis of Stenotrophomonas maltophilia reveals an organism heavily shielded by drug resistance determinants. Genome Biol 9:R74. doi: 10.1186/gb-2008-9-4-r74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Doménech-Sánchez A, Martínez-Martínez L, Hernández-Allés S, del Carmen Conejo M, Pascual A, Tomás JM, Albertí S, Benedí VJ. 2003. Role of Klebsiella pneumoniae OmpK35 porin in antimicrobial resistance. Antimicrob Agents Chemother 47:3332–3335. doi: 10.1128/AAC.47.10.3332-3335.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Moya-Torres A, Mulvey MR, Kumar A, Oresnik IJ, Brassinga AKCB. 2014. The lack of OmpF, but not OmpC, contributes to increased antibiotic resistance in Serratia marcescens. Microbiology (Reading) 160:1882–1892. doi: 10.1099/mic.0.081166-0. [DOI] [PubMed] [Google Scholar]
- 41.Wolter DJ, Hanson ND, Lister PD. 2004. Insertional inactivation of oprD in clinical isolates of Pseudomonas aeruginosa leading to carbapenem resistance. FEMS Microbiol Lett 236:137–143. doi: 10.1016/j.femsle.2004.05.039. [DOI] [PubMed] [Google Scholar]
- 42.Woodruff WA, Hancock RE. 1988. Construction and characterization of Pseudomonas aeruginosa protein F-deficient mutants after in vitro and in vivo insertion mutagenesis of the cloned gene. J Bacteriol 170:2592–2598. doi: 10.1128/jb.170.6.2592-2598.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Masi M, Winterhalter M, Pagès JM. 2019. Outer membrane porins. Subcell Biochem 92:79–123. doi: 10.1007/978-3-030-18768-2_4. [DOI] [PubMed] [Google Scholar]
- 44.Smani Y, Fàbrega A, Roca I, Sánchez-Encinales V, Vila J, Pachón J. 2014. Role of OmpA in the multidrug resistance phenotype of Acinetobacter baumannii. Antimicrob Agents Chemother 58:1806–1808. doi: 10.1128/AAC.02101-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Ross CL, Thomason KS, Koehler TM. 2009. An extracytoplasmic function sigma factor controls beta-lactamase gene expression in Bacillus anthracis and other Bacillus cereus group species. J Bacteriol 191:6683–6693. doi: 10.1128/JB.00691-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Ho TD, Nauta KM, Müh U, Ellermeier CD. 2019. Activation of the extracytoplasmic function sigma factor sigma(P) by beta-lactam in Bacillus thuringiensis requires the site-2 protease RasP. mSphere 4:e00511-19. doi: 10.1128/mSphere.00511-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Luo Y, Helmann JD. 2012. Analysis of the role of Bacillus subtilis σ(M) in β-lactam resistance reveals an essential role for c-di-AMP in peptidoglycan hemostasis. Mol Microbiol 83:623–639. doi: 10.1111/j.1365-2958.2011.07953.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Lee YH, Helmann JD. 2014. Mutations in the primary sigma factor sigmaA and termination factor rho that reduce susceptibility to cell wall antibiotics. J Bacteriol 196:3700–3711. doi: 10.1128/JB.02022-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Morikawa K, Maruyama A, Inose Y, Higashide M, Hayashi H, Ohta T. 2001. Overexpression of sigma factor, sigma(B), urges Staphylococcus aureus to thicken the cell wall and to resist beta-lactams. Biochem Biophys Res Commun 288:385–389. doi: 10.1006/bbrc.2001.5774. [DOI] [PubMed] [Google Scholar]
- 50.Balasubramanian D, Kong KF, Jayawardena SR, Leal SM, Sautter RT, Mathee K. 2011. Co-regulation of β-lactam resistance, alginate production and quorum sensing in Pseudomonas aeruginosa. J Med Microbiol 60:147–156. doi: 10.1099/jmm.0.021600-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Tettmann B, Dötsch A, Armant O, Fjell CD, Overhage J. 2014. Knockout of extracytoplasmic function sigma factor ECF-10 affects stress resistance and biofilm formation in Pseudomonas putida KT2440. Appl Environ Microbiol 80:4911–4919. doi: 10.1128/AEM.01291-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Pandey P, Dubey AP, Mishra S, Singh VS, Singh C, Tripathi AK. 2022. β-lactam resistance in Azospirillum baldaniorum Sp245 is mediated by lytic transglycosylase and β-lactamase and regulated by a cascade of RpoE7→RpoH3 sigma factors. J Bacteriol 204:e0001022. doi: 10.1128/jb.00010-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Helmann JD. 1999. Anti-sigma factors. Curr Opin Microbiol 2:135–141. doi: 10.1016/S1369-5274(99)80024-1. [DOI] [PubMed] [Google Scholar]
- 54.Ho TD, Ellermeier CD. 2012. Extra cytoplasmic function σ factor activation. Curr Opin Microbiol 15:182–188. doi: 10.1016/j.mib.2012.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Li LH, Zhang MS, Wu CJ, Lin YT, Yang TC. 2019. Overexpression of SmeGH contributes to the acquired MDR of Stenotrophomonas maltophilia. J Antimicrob Chemother 74:2225–2229. doi: 10.1093/jac/dkz200. [DOI] [PubMed] [Google Scholar]
- 56.Yang TC, Huang YW, Hu RM, Huang SC, Lin YT. 2009. AmpDI is involved in expression of the chromosomal L1 and L2 beta-lactamases of Stenotrophomonas maltophilia. Antimicrob Agents Chemother 53:2902–2907. doi: 10.1128/AAC.01513-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Livak KJ, Schmittgen TD. 2001. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) method. Methods 25:402–408. doi: 10.1006/meth.2001.1262. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Fig. S1 to S3 and Tables S1 to S4, S6, and S7. Download spectrum.02797-22-s0001.pdf, PDF file, 5.3 MB (5.4MB, pdf)
Table S5. Download spectrum.02797-22-s0002.xlsx, XLSX file, 0.2 MB (160.1KB, xlsx)
Data Availability Statement
The RNA-seq data have been deposited in GenBank BioSample accessions SAMN25290492 for S. maltophilia KJ and SAMN30672914 for KJΔOmpA299-356.